Enhanced Olivine Carbonation within a Basalt as Compared to Single

Apr 16, 2014 - Institut des Sciences de la Terre, UMR 5275 CNRS, Université de Grenoble 1, 1381 rue de la Piscine, 38400 Saint Martin d,Hères,. Fran...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/est

Enhanced Olivine Carbonation within a Basalt as Compared to Single-Phase Experiments: Reevaluating the Potential of CO2 Mineral Sequestration Olivier Sissmann,*,†,‡,§ Fabrice Brunet,∥ Isabelle Martinez,† François Guyot,⊥ Anne Verlaguet,#,¶ Yves Pinquier,‡ and Damien Daval∇ †

Institut de Physique du Globe de Paris, Sorbonne Paris Cité, Université Paris Diderot, UMR 7154 CNRS, 1 rue Jussieu, F-75005 Paris, France ‡ Laboratoire de Géologie, UMR 8538 CNRS, École Normale Supérieure, 24 rue Lhomond, 75005 Paris, France § IFP Energies Nouvelles, 1 et 4 avenue de Bois-Préau, 92852 Rueil-Malmaison, France ∥ Institut des Sciences de la Terre, UMR 5275 CNRS, Université de Grenoble 1, 1381 rue de la Piscine, 38400 Saint Martin d’Hères, France ⊥ Institut de Minéralogie, de Physique des Matériaux et de Cosmochimie, UMR 7590 CNRS, Museum National d’Histoire Naturelle/Université Pierre et Marie Curie, 4 place Jussieu, 75005 Paris, France # UPMC Univ. Paris 06, UMR 7193, ISTeP, F-75005 Paris, France ∇ Laboratoire d’Hydrologie et de Géochimie de Strasbourg, Université de Strasbourg/EOST CNRS UMR 7517, 1 rue Blessig, 67084 Strasbourg, France ¶ CNRS, UMR 7193, ISTeP, F-75005 Paris, France S Supporting Information *

ABSTRACT: Batch experiments were conducted in water at 150 °C and PCO2 = 280 bar on a Mg-rich tholeiitic basalt (9.3 wt % MgO and 12.2 wt % CaO) composed of olivine, Ti-magnetite, plagioclase, and clinopyroxene. After 45 days of reaction, 56 wt % of the initial MgO had reacted with CO2 to form Fe-bearing magnesite, (Mg0.8Fe0.2)CO3, along with minor calcium carbonates. The substantial decrease in olivine content upon carbonation supports the idea that ferroan magnesite formation mainly follows from olivine dissolution. In contrast, in experiments performed under similar run durations and P/T conditions with a San Carlos olivine separate (47.8 wt % MgO) of similar grain size, only 5 wt % of the initial MgO content reacted to form Fe-bearing magnesite. The overall carbonation kinetics of the basalt was enhanced by a factor of ca. 40. This could be explained by differences in the chemical and textural properties of the secondary silica layer that covers reacted olivine grains in both types of sample. Consequently, laboratory data obtained on olivine separates might yield a conservative estimate of the true carbonation potential of olivine-bearing basaltic rocks.



INTRODUCTION Basalts are mafic rocks containing minerals rich in divalent, solid-carbonate-forming cations (Ca, Mg, Fe). Therefore, they are good candidates for CO2 sequestration.1−4 In that respect, basaltic sites for pilot CO2 injections have been selected in the Columbia River formation5 and on the Stapafell Mountain in Hellisheidi, Iceland.4,6 Among the basalt-forming minerals, olivine, (Mg,Fe)2SiO4, shows the highest carbonation potential since it combines a high (Mg,Fe) content and rapid dissolution kinetics.7,8 However, previous batch carbonation studies on separated olivine grains have emphasized the deleterious role played by secondary phases, such as surface layers rich in amorphous silica, SiO2(am), in the dissolution rate of the © 2014 American Chemical Society

parent mineral and the transport of reactants from and to the reactive surface (e.g., see refs 9−12). It is therefore to be expected that the carbonation efficiency of a polymineral rock such as a basalt will also depend on the nature and texture of those secondary layers (mainly on their chemical composition and permeability). In the present study, the carbonation efficiency of olivine in a Stapafell basalt sample was compared with that of San Carlos Received: Revised: Accepted: Published: 5512

December 10, 2013 April 16, 2014 April 16, 2014 April 16, 2014 dx.doi.org/10.1021/es405508a | Environ. Sci. Technol. 2014, 48, 5512−5519

Environmental Science & Technology

Article

Table 1. Solution Composition and Corresponding Thermodynamic Calculations for the 10 Experiments Carried out in the Present Studya ΔGr (kJ/mol) sample basalt

San Carlos olivine

t (days)

[Mg(aq)] (mmol/L)

[SiO2(am)] (mmol/L)

pH0

pHf

forsterite

2 5 10 20 30 45 2 5 10 30

1.4 7.8 5 5.2 4.9 5.1 9.05 9.71 5.88 5.33

5.39 15.3 11.8 11.7 12 11.9 4.96 5.24 8.35 8.28

3.24 3.24 3.24 3.24 3.24 3.24 3.24 3.24 3.24 3.24

3.89 4.54 4.37 4.41 4.37 4.42 4.59 4.62 4.43 4.4

−78.28 −45.75 −51.74 −53.21 −54.10 −53.47 −47.37 −46.08 −52.40 −54.03

magnesite SiO2(am) −11.02 3.41 0.87 0.15 −0.34 −0.01 4.56 5.13 1.15 0.35

−2.26 1.41 0.50 0.47 0.56 0.53 −2.55 −2.36 −0.72 −0.75

amount in powder (wt %) talc

(Mg,Fe)CO3

CaCO3

carb. MgO

−54.93 3.05 −8.23 −10.52 −11.62 −10.75 −9.30 −6.88 −12.26 −14.78

0.08 2.30 4.80 7.80 11.80 14.80 nd 0.10 2.41 4.62

nd nd 0.01 0.01 0.01 0.03 nd nd nd nd

0.30 8.65 18.06 29.34 44.39 55.67 nd 0.10 2.39 4.59

The first two columns indicate the type of sample and the duration of the experiment. In the next two columns are listed the concentrations of Mg and Si in the fluid (mmol/L) measured by ICP-AES at the end of each experiment (with uncertainties of ±5%). Ca and Na are not reported, as quenching effects affected the measured values; Al, Fe, Mn, and Ni were below detection limit. The initial and final pH calculated with CHESS are reported in the subsequent two columns. In the four following columns are listed the Gibbs free energies of the dissolution reactions of forsterite, magnesite, amorphous silica, and talc (kJ/mol). For the sake of simplicity, only the concentrations of Mg and Si were considered in the thermodynamic calculations, as they are only negligibly affected by the concentration of other cations. The last three columns indicate the proportions of (Mg,Fe)CO3 and CaCO3 (wt %) in the powders after different time durations, and the weight percentage of initial MgO that carbonated (carb. MgO). a

olivine starting material contained no other phases aside from minor quantities of iron oxide, which were removed manually under the binoculars. As determined from electron microprobe analyses (Table S1), the San Carlos olivine (Fo88) was slightly poorer in iron than the basalt olivine (Fo85). The specific surface area of the basalt powder measured by the Brunauer−Emmett−Teller (BET) method was 0.40 m2·g−1, while that of San Carlos olivine was 0.13 m2·g−1; the measurements were conducted with an accuracy of 10%. Importantly, we verified using SEM that the grain size of olivine in the basalt powder (averaging 41 ± 12 μm; Figure S1) was similar to that of the San Carlos olivine, so the difference between the BET specific surface areas of those two samples cannot be ascribed to differences in olivine grain size. Instead, the higher BET surface area of the basalt may be attributed to the multiphasic nature of some of the sieved particles (mostly, plagioclase/pyroxene and occasionally olivine−pyroxene aggregates), which hosted additional grain boundaries compared with the sieved San Carlos olivine grains (see, e.g., ref 16 for the detailed reasoning). Since the largest phenocrystals observed within the powder were olivine, which represents only 12 wt % of the basalt, it is likely that most of this BET surface area difference is due to the smaller, more abundant plagioclase and pyroxene crystals (see Figure S1e). Experimental Conditions and Protocol. The experiments were conducted at T = 150 °C and PCO2 = 280 bar in ultrapure deionized water. Either basalt or olivine powder (5 g) and 50 mL of water were introduced into a Teflon liner fitted inside a 100 mL Ti autoclave. The reactor was then sealed by compressing the head of the autoclave on a circular Viton gasket. At the end of each experiment, the temperature was first decreased to ambient conditions, after which CO2 was flushed out of the autoclave and the reactor was opened. The fluid was then extracted with a syringe and immediately diluted 20-fold in ultrapure water and acidified with HNO3 (2 vol %). It was subsequently analyzed for major cations on a Varian 720 ES inductively coupled plasma atomic emission spectrometry (ICP-AES) instrument (ISTerre, Grenoble, France) with

olivine on the basis of experiments performed on powdered samples at 150 °C under CO2 at a pressure of 280 bar. Although the temperature studied is in the higher range of what could be considered acceptable for in situ geological storage,13 it allows for faster dissolution kinetics of the primary silicate assemblage. Moreover, working in this P/T range is relevant for ex situ CO2 sequestration purposes (e.g., see ref 14). In an attempt to unravel the differences of olivine reactivity in these two series of experiments, the nature, texture, and chemical composition of interfacial secondary phases produced during the course of the dissolution−carbonation process were characterized by electron microscopy down to the nanometer scale.



MATERIALS AND METHODS

Starting Materials. This study was conducted on a midocean ridge basalt (MORB) sample from the Stapafell Mountain, Iceland (the whole rock and mineral compositions are listed in Table S1 in the Supporting Information). Four main crystalline phases were identified using scanning electron microscopy (SEM) (Figure S1 in the Supporting Information) and X-ray diffraction (XRD) (Figure S2a). Normative mineral calculations indicated that the sample had weight proportions of 44 wt % plagioclase, 38 wt % clinopyroxene (augitic pyroxene), 12 wt % olivine, and 6 wt % iron oxide (titanomagnetite), in agreement with values previously reported for samples from this site.15 Variable amounts of glass, up to ∼4 wt %, were previously reported,15 but no glass phase was identified by SEM and optical microscopy in the specific samples studied here. The sample was crushed and sieved to a 33−80 μm grain size fraction and subsequently cleaned in ethanol several times to remove fine particles. The powder was then rinsed with ultrapure deionized water and dried overnight at 50 °C. SEM confirmed that most of the fine particles were removed from the surface of the larger grains. Similar preparation was conducted on a batch of gem-quality San Carlos olivine, (Mg0.88Fe0.12)2SiO4 (Table S1), to obtain a powder in the same grain size range (33−80 μm). This separate 5513

dx.doi.org/10.1021/es405508a | Environ. Sci. Technol. 2014, 48, 5512−5519

Environmental Science & Technology

Article

analytical uncertainties below 5%. Finally, the solid reaction products were dried overnight at 50 °C. A total of six different experiments for the basalt and four for the olivine were conducted under these conditions with run durations varying from 2 to 45 days (Table 1). Sample Characterization. XRD patterns of the solid products were collected with a Rigaku ultraX18HFCE Bragg− Brentano diffractometer equipped with a rotating copper anode at ENS (Paris, France). The X-ray beam was generated at 300 mA and 50 kV, and scans were performed over the 2θ range from 5 to 90° with 0.03° per step and a total accumulation time of ∼50 min. A portion of the reaction products was fixed on an adhesive carbon holder, gold-coated, and studied using a fieldemission gun (FEG) scanning electron microscope (Zeiss Ultra, ENS Paris) equipped with an energy-dispersive silicon drift X-ray (EDS) detector (X-Max, Oxford Instruments). Standard electron images were collected under a 15 kV voltage at a working distance of 8 mm. In addition, settings of 5 kV and a 3 mm working distance were used to produce high-quality secondary electron (SE) images of mineral surfaces. Focused ion beam (FIB) milling of the reaction products using a HELIOS 600 Nanolab dual-beam FIB (CP2M, Marseille, France) was also performed to obtain electrontransparent ultrathin sections. The samples were carbon-coated beforehand in order to minimize the production of surface artifacts due to Pt and Ga beam interactions during the milling procedure. Once extracted from the samples, the FIB cross sections were observed by transmission electron microscopy (TEM) with a JEOL 2100F FEG operated at 200 kV (IMPMC, Université Paris-VI). The amounts of produced carbonates were quantified using a classical technique of CO 2 extraction based on their decomposition in orthophosphoric acid.17 The reaction product was placed in a Pyrex tube on a vacuum line containing an isolated volume of orthophosphoric acid. After a vacuum of 5 days), we were able to show that the olivine in both the San Carlos and basalt samples has slower dissolution kinetics than expected on the basis of kinetic simulations using the whole range of dissolution rate laws that can be retrieved from the literature (Figure 5 and Table S2; see ref 10 for details). As mentioned above, a likely effect that could explain the gap between the experimental data and the simulations may be sought in the role of secondary phases on olivine dissolution (i.e., transport limitation and/or surface passivation), which was not taken into account by the kinetic model here provided by CHESS. For the sake of example, a satisfactory fit of the data could be obtained by lowering the actual specific surface area (SSA) of olivine by a factor of 200 for the San Carlos olivine experiments and by a factor of 5 for the basalt experiments. Hence, a prominent result of the present study is that the olivine carbonation kinetics was at least 40 times slower in the mineral separate (San Carlos olivine) than in the tholeiitic basalt (Figure 5). Even though the basalt has an SSA that is 3 times superior to that of San Carlos olivine (mostly as a result of the smaller and more abundant other phases in the basalt grains, but to which olivine grain boundary surfaces and internal pores may nonetheless have contributed), this is not enough to explain this very large difference in reactivity. Numerous experimental studies9−11,27,28 have shown that olivine dissolution, and thus its carbonation, is impeded by the formation of a protective secondary amorphous Si-rich layer. A 5517

dx.doi.org/10.1021/es405508a | Environ. Sci. Technol. 2014, 48, 5512−5519

Environmental Science & Technology

Article

phases in the starting material. This material is available free of charge via the Internet at http://pubs.acs.org.

polymerization process, as has been suggested for the incorporation of Zr and Ca into silica gels formed on altered nuclear glass,34,35 and could then account for the apparent porosity of the layer and its potential permeability to the transport of Mg and Fe toward the bulk solution. Previous studies have also emphasized that reducing conditions can alter the dissolution rate of olivine,36,37 potentially because of underlying modifications of the passivating properties of amorphous silica around olivine.12,25 Two recent studies12,25 indirectly showed that in experimental setups such as the one used here, the redox conditions are partly controlled by iron in the solid sample (valency, content, and mineralogy). More specifically, their results suggested that under the initial redox conditions imposed by water having equilibrated with air, the iron contained in olivine would first precipitate as insoluble Fe(III) and be incorporated within the interfacial silica layer, dramatically lowering the olivine dissolution rate. As the redox process evolves with the progressive consumption of O2 in the chemically closed reactor, Fe(III) becomes unstable and is reduced to Fe(II), which is far more soluble. Consequently, the silica layer is partly destabilized by release of Fe(II) into the aqueous medium, allowing dissolution of the primary silicates to proceed faster. It is therefore likely that in our powder experiments performed on the titanomagnetite-bearing Stapafell mountain basalt and on single-phase San Carlos olivine, the redox conditions followed different evolutions, at least transiently (see Figure S8), and therefore are another chemical parameter that may have contributed to the observed differences in the overall carbonation yields. Implications for CO2 Sequestration via Basalt Carbonation. Despite the low abundance of olivine within basaltic rocks and the severe passivation of olivine surfaces encountered in carbonation experiments on the single-phase mineral separate, our study suggests that basalts may be considered as an appealing target for carbonation purposes, as a yield of 77 g of stored CO2 per kilogram of basalt powder was reached after 45 days of reaction at 150 °C and over half of the initial magnesium content of the rock was carbonated. By combining measurements of carbonation yields and the characterization of secondary phases down to the nanometer scale, we suggest that the incorporation of Al within the interfacial amorphous silica layer covering olivine may enhance its permeability and thus favor further olivine dissolution. Alternatively, favorable redox conditions may have limited the passivation ability of the amorphous silica layers.12,25 Although a direct upscaling of the present results to natural settings is precluded because of the complexity of the process at the rock scale (e.g., fluid circulation in fractured rocks and fluid heterogeneities), which could partly be addressed by running percolation experiments on rock samples (e.g., see ref 27), the present results nonetheless undoubtedly revive the carbonation potential of olivine-bearing basaltic rocks.





AUTHOR INFORMATION

Corresponding Author

*Phone: +33 1 47 52 50 09; e-mail: [email protected]. Notes

The authors declare no competing financial interest. This is IPGP contribution no. 3519.



ACKNOWLEDGMENTS The authors thank Damien Deldicque for his help with SEM and XRD analysis, Christian Dominici at the CP2M (Marseilles) for help with FIB milling, Carine Chaduteau for help with carbonate quantification, Jean-Michel Guigner at IMPMC (Paris) for help with TEM analysis, Frederique Metzelard for her availability and help in solving administrative problems in the lab, and Delphine Tisserand, Lionel Rosseto, and Sarah Bureau for help with ICP analysis. Christian Chopin, ̈ Pierre Agrinier, Benedicte Menez, and Aicha El-Khamlichi are warmly acknowledged for continuous support and stimulating exchanges. The committee members of the IPGP-ADEMETotal-Schlumberger Research Group on CO2 sequestration and of the ANR CO2FIX (ANR-08-PCO2-003-03) are thanked for fruitful discussions and funding. Finally, the thorough reviews and constructive comments by three anonymous reviewers and the editor were also much appreciated and helped improve the manuscript.



REFERENCES

(1) Dessert, C.; Dupré, B.; Gaillardet, J.; François, L. M.; Allègre, C. J. Basalt weathering laws and the impact of basalt weathering on the global carbon cycle. Chem. Geol. 2003, 202, 257−273. (2) McGrail, B. P.; Schaef, H. T.; Ho, A. M.; Chien, Y.-J.; Dooley, J. J.; Davidson, C. L. Potential for carbon dioxide sequestration in flood basalts. J. Geophys. Res.: Solid Earth 2006, 111, No. B12201. (3) Kelemen, P. B.; Matter, J. In situ carbonation of peridotite for CO2 storage. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 17295−17300. (4) Gislason, S. R.; Wolff-Boenisch, D.; Stefansson, A.; Oelkers, E. H.; Gunnlaugsson, E.; Sigurdardottir, H.; Sigfusson, B.; Broecker, W. S.; Matter, J. M.; Stute, M.; Axelsson, G.; Fridriksson, T. Mineral sequestration of carbon dioxide in basalt: A pre-injection overview of the CarbFix project. Int. J. Greenhouse Gas Control 2010, 4, 537−545. (5) Schaef, H. T.; McGrail, B. P. Dissolution of Columbia River basalt under mildly acidic conditions as a function of temperature: Experimental results relevant to the geological sequestration of carbon dioxide. Appl. Geochem. 2009, 24, 980−987. (6) Alfredsson, H. A.; Oelkers, E. H.; Hardarsson, B. S.; Franzson, H.; Gunnlaugsson, E.; Gislason, S. R. The geology and water chemistry of the Hellisheidi, SW-Iceland carbon storage site. Int. J. Greenhouse Gas Control 2013, 12, 399−418. (7) Guyot, F.; Daval, D.; Dupraz, S.; Martinez, I.; Menez, B.; Sissmann, O. CO2 geological storage: The environmental mineralogy perspective. C. R. Geosci. 2011, 343, 246−259. (8) Garcia, B.; Beaumont, V.; Perfetti, E.; Rouchon, V.; Blanchet, D.; Oger, P.; Dromart, G.; Huc, A. Y.; Haeseler, F. Experiments and geochemical modelling of CO2 sequestration by olivine: Potential, quantification. Appl. Geochem. 2010, 25, 1383−1396. (9) Bearat, H.; McKelvy, M. J.; Chizmeshya, A. V. G.; Gormley, D.; Nunez, R.; Carpenter, R. W.; Squires, K.; Wolf, G. H. Carbon sequestration via aqueous olivine mineral carbonation: Role of passivating layer formation. Environ. Sci. Technol. 2006, 40, 4802− 4808. (10) Daval, D.; Sissmann, O.; Menguy, N.; Saldi, G. D.; Guyot, F.; Martinez, I.; Corvisier, J.; Garcia, B.; Machouk, I.; Knauss, K. G.; Hellmann, R. Influence of amorphous silica layer formation on the

ASSOCIATED CONTENT

S Supporting Information *

Information about the morphology, size, and chemical composition of the minerals in the starting materials for both sets of experiments as well as of the secondary phases forming during these experiments; detailed carbonation results of the experiments; a numerical simulation for modeling the evolution of Eh during both sets of experiments; and calculations at relevant P/T conditions for the dissolution of the mineral 5518

dx.doi.org/10.1021/es405508a | Environ. Sci. Technol. 2014, 48, 5512−5519

Environmental Science & Technology

Article

dissolution rate of olivine at 90 °C and elevated pCO2. Chem. Geol. 2011, 284, 193−209. (11) Wang, F.; Giammar, D. E. Forsterite dissolution in saline water at elevated temperature and high CO2 pressure. Environ. Sci. Technol. 2013, 47, 168−173. (12) Saldi, G. D.; Daval, D.; Morvan, G.; Knauss, K. G. The role of Fe and redox conditions in olivine carbonation rates: An experimental study of the rate limiting reactions at 90 and 150 °C in open and closed systems. Geochim. Cosmochim. Acta 2013, 118, 157−183. (13) Bachu, S. Sequestration of CO2 in geological media: Criteria and approach for site selection in response to climate change. Energy Convers. Manage. 2000, 41, 953−970. (14) Gerdemann, S. J.; O’Connor, W. K.; Dahlin, D. C.; Penner, L. R.; Rush, H. Ex situ aqueous mineral carbonation. Environ. Sci. Technol. 2007, 41, 2587−2593. (15) Gudbrandsson, S.; Wolff-Boenisch, D.; Gislason, S. R.; Oelkers, E. H. An experimental study of crystalline basalt dissolution from 2 < pH < 11 and temperatures from 5 to 75 °C. Geochim. Cosmochim. Acta 2011, 75, 5496−5509. (16) Daval, D.; Hellmann, R.; Martinez, I.; Gangloff, S.; Guyot, F. Lizardite serpentine dissolution kinetics as a function of pH and temperature, including effects of elevated pCO2. Chem. Geol. 2013, 351, 245−256. (17) McCrea, J. M. On the isotopic chemistry of carbonates and a paleotemperature scale. J. Chem. Phys. 1950, 18, 849−857. (18) Barnes, C. S.; Koretsky, M. D. ThermoSolver Software; John Wiley and Sons: Hoboken, NJ, 2004 (accompanies the textbook Engineering and Chemical Thermodynamics, 2nd ed., by M. D. Koretsky). (19) van der Lee, J.; De Windt, L. CHESS Tutorial and Cookbook, Updated for Version 3.0; Manual no. LHM/RD/02/13; Ecole des Mines de Paris: Fontainebleau, France, 2002; p 116. (20) Wolery, T. J.; Daveler, S. A. EQ6, a Computer Program for Reaction Path Modeling of Aqueous Geochemical Systems: Theoretical Manual, User’s Guide, and Related Documentation, version 7.0; Lawrence Livermore National Laboratory: Livermore, CA, 1992; p 349. (21) Sorai, M.; Sasaki, M. Dissolution kinetics of anorthite in a supercritical CO2−water system. Am. Mineral. 2010, 95, 853−862. (22) Oelkers, E. H.; Schott, J. Experimental study of anorthite dissolution and the relative mechanism of feldspar hydrolysis. Geochim. Cosmochim. Acta 1995, 59, 5039−5053. (23) Stillings, L. L.; Brantley, S. L. Feldspar dissolution at 25 °C and pH 3: Reaction stoichiometry and the effect of cations. Geochim. Cosmochim. Acta 1995, 59, 1483−1496. (24) Rosenbauer, R. J.; Thomas, B.; Bischoff, J. L.; Palandri, J. Carbon sequestration via reaction with basaltic rocks: Geochemical modeling and experimental results. Geochim. Cosmochim. Acta 2012, 89, 116−133. (25) Sissmann, O.; Daval, D.; Brunet, F.; Guyot, F.; Verlaguet, A.; Pinquier, Y.; Findling, N.; Martinez, I. The deleterious effect of secondary phases on olivine carbonation yield: Insight from timeresolved aqueous-fluid sampling and FIB-TEM characterization. Chem. Geol. 2013, 357, 186−202. (26) Saldi, G. D.; Schott, J.; Pokrovsky, O. S.; Gautier, Q.; Oelkers, E. H. An experimental study of magnesite precipitation rates at neutral to alkaline conditions and 100−200 °C as a function of pH, aqueous solution composition and chemical affinity. Geochim. Cosmochim. Acta 2012, 83, 93−109. (27) Andreani, M.; Luquot, L.; Gouze, P.; Godard, M.; Hoise, E.; Gibert, B. Experimental study of carbon sequestration reactions controlled by the percolation of CO2-rich brine through peridotites. Environ. Sci. Technol. 2009, 43, 1226−1231. (28) Johnson, N. C.; Thomas, B.; Maher, K.; Rosenbauer, R. J.; Bird, D.; Brown, G. E., Jr. Olivine dissolution and carbonation under conditions relevant for in situ carbon storage. Chem. Geol. 2014, 373, 93−105.

(29) Andreani, M.; Daniel, I.; Pollet-Villard, M. Aluminum speeds up the hydrothermal alteration of olivine. Am. Mineral. 2013, 98, 1738− 1744. (30) King, H. E.; Plumper, O.; Geisler, T.; Putnis, A. Experimental investigations into the silicification of olivine: Implications for the reaction mechanism and acid neutralization. Am. Mineral. 2011, 96, 1503−1511. (31) Hellmann, R.; Wirth, R.; Daval, D.; Barnes, J.-P.; Penisson, J.M.; Tisserand, D.; Epicier, T.; Florin, B.; Hervig, R. L. Unifying natural and laboratory chemical weathering with interfacial dissolution− reprecipitation: A study based on the nanometer-scale chemistry of fluid−silicate interfaces. Chem. Geol. 2012, 294−295, 203−216. (32) King, H. E.; Plumper, O.; Putnis, A. Effect of secondary phase formation on the carbonation of olivine. Environ. Sci. Technol. 2010, 44, 6503−6509. (33) Morris, P. M.; Wogelius, R. A. Phthalic acid complexation and the dissolution of forsteritic glass studied via in situ FTIR and X-ray scattering. Geochim. Cosmochim. Acta 2008, 72, 1970−1985. (34) Cailleteau, C.; Angeli, F.; Devreux, F.; Gin, S.; Jestin, J.; Jollivet, P.; Spalla, O. Insight into silicate-glass corrosion mechanisms. Nat. Mater. 2008, 7, 978−983. (35) Jollivet, P.; Gin, S.; Schumacher, S. Forward dissolution rate of silicate glasses of nuclear interest in clay-equilibrated groundwater. Chem. Geol. 2012, 330−331, 207−217. (36) Wogelius, R. A.; Walther, J. V. Olivine dissolution kinetics at near-surface conditions. Chem. Geol. 1992, 97, 101−112. (37) Wogelius, R. A.; Walther, J. V. Olivine dissolution at 25 °C: Effects of pH, CO2, and organic acids. Geochim. Cosmochim. Acta 1991, 55, 943−954. (38) Hänchen, M.; Prigiobbe, V.; Storti, G.; Seward, T. M.; Mazzotti, M. Dissolution kinetics of fosteritic olivine at 90−150 °C, including effects of the presence of CO2. Geochim. Cosmochim. Acta 2006, 70, 4403−4416. (39) Rosso, J. J.; Rimstidt, D. J. A high resolution study of forsterite dissolution rates. Geochim. Cosmochim. Acta 2000, 64, 797−811.

5519

dx.doi.org/10.1021/es405508a | Environ. Sci. Technol. 2014, 48, 5512−5519